You are on page 1of 16

polymers

Article
Effect of Infill Density in FDM 3D Printing on Low-Cycle Stress
of Bamboo-Filled PLA-Based Material
Miroslav Müller 1 , Petr Jirků 1 , Vladimír Šleger 2 , Rajesh Kumar Mishra 1, * , Monika Hromasová 3
and Jan Novotný 4

1 Department of Material Science and Manufacturing Technology, Faculty of Engineering, Czech University of
Life Sciences Prague, Kamycka 129, 165 00 Prague, Czech Republic
2 Department of Mechanical Engineering, Faculty of Engineering, Czech University of Life Sciences Prague,
Kamycka 129, 165 00 Prague, Czech Republic
3 Department of Electrical Engineering and Automation, Faculty of Engineering, Czech University of Life
Sciences Prague, Kamycka 129, 165 00 Prague, Czech Republic
4 Faculty of Mechanical Engineering, J. E. Purkyne University in Usti nad Labem, Pasteurova 3334/7,
400 01 Usti nad Labem, Czech Republic
* Correspondence: mishrar@tf.czu.cz

Abstract: In this paper, the fatigue behavior of polylactic acid (PLA) material with bamboo filler
printed by 3D additive printing using fused deposition modelling (FDM) technology at different infill
densities and print nozzle diameters is investigated. The mechanical test results are supported by the
findings from SEM image analysis. The fatigue behavior was tested at four consecutive 250 cycles at
loads ranging from 5 to 20, 30, 40, and 50% based on the limits found in the static tensile test. The
results of the static tensile and low-cycle fatigue tests confirmed significant effects of infill density
of 60%, 80%, and 100% on the tensile strength of the tested specimens. In particular, the research
results show a significant effect of infill density on the fatigue properties of the tested materials. The
Citation: Müller, M.; Jirků, P.; Šleger,
influence of cyclic tests resulted in the strengthening of the tested material, and at the same time,
V.; Mishra, R.K.; Hromasová, M.;
its viscoelastic behavior was manifested. SEM analysis of the fracture surface confirmed a good
Novotný, J. Effect of Infill Density in
interaction between the PLA matrix and the bamboo-based filler using nozzle diameters of 0.4 and
FDM 3D Printing on Low-Cycle
0.6 mm and infill densities of 60%, 80%, and 100%. Low-cycle testing showed no reductions in the
Stress of Bamboo-Filled PLA-Based
Material. Polymers 2022, 14, 4930.
mechanical properties and fatigue lives of the 3D printed samples.
https://doi.org/10.3390/
polym14224930 Keywords: 3D printing; PLA polymer; bamboo; low-cycle test; infill density; mechanical properties; SEM

Academic Editors: Mario Bragaglia


and Francesca Nanni

Received: 14 October 2022 1. Introduction


Accepted: 11 November 2022 The current trend in the field of polymeric materials and related composites is towards
Published: 15 November 2022 the use of biological material, not only in the field of processing by injection moulding, vacuum
Publisher’s Note: MDPI stays neutral infusion, and bonding technology, but also in the field of 3D printing. The use of biological
with regard to jurisdictional claims in material in the field of vacuum infusion and bonding technology is mainly related to the filler
published maps and institutional affil- and the matrix, which are usually made of a synthetic resin. In the field of polymer injection
iations. moulding and 3D printing, both the biological matrix and the filler are used [1].
New materials, known as biocomposites, are emerging and belong to a group of
materials with significant research potential. Among them, PLA is one of the most widely
used biodegradable filament materials in the field of additive technologies [2]. However,
Copyright: © 2022 by the authors. its disadvantages include brittleness, low toughness, and flexibility [3]. Biocomposite
Licensee MDPI, Basel, Switzerland.
filaments include materials with biodegradable polymeric matrices and bio-based fillers.
This article is an open access article
The fillers can be in the form of particles or fibres. Particulate fillers are mainly used for
distributed under the terms and
additive technologies due to the nozzle deposition of the material. The filler content is
conditions of the Creative Commons
usually up to 50% by volume when using natural materials [4]. The literature suggests
Attribution (CC BY) license (https://
a wide variation in the natural fillers added to the printing material used in 3D printing,
creativecommons.org/licenses/by/
especially for sustainability reasons [5–7]. The most widely used thermoplastic polymer
4.0/).

Polymers 2022, 14, 4930. https://doi.org/10.3390/polym14224930 https://www.mdpi.com/journal/polymers


Polymers 2022, 14, 4930 2 of 16

in biocomposites in 3D printing is PLA. Fibre manufacturers use many types of cellulose


or other natural fibres or particles as filler. The fibres used include: bamboo, birch, cherry,
cedar, coconut, cork, ebony, olive, pine, or willow [8]. Research results show that bio-based
fillers in combination with additive technology reduce production costs [9,10]. Examples
are discoveries reported on kenaf particles, wood flour, etc. [9,10] However, changes in the
mechanical properties of the filament must also be considered, which are highly dependent
on the bio-based filler used, its treatment, and the interaction between the filler and the
matrix [7,11–15]. The natural filler provides considerable variability in the microstructures
and compositions of the different species [12].
There are many different technologies in the field of additive manufacturing or 3D
printing. The most common and very popular method among users is additive FDM (fused
deposition modelling) technology, which works with printing filament. This filament can be
based on primary, secondary, or composite materials using different fillers [1,6]. The compo-
nents printed using additive FDM technology typically have poorer mechanical behaviours
than the same materials processed by injection moulding, stamping, or extrusion [16,17].
The quality of 3D-printed components, and consequently, their mechanical properties,
are influenced by various parameters affecting interlayer fusion, porosity, and swelling
caused by natural fibres and particles, etc. [17,18]. On the other hand, additive technology
offers designers more freedom and the possibility of addressing specific problems on an
object, for example, by increasing the infill density [16,17]. Additive manufacturing in-
volves the production of material in layers and is highly dependent on the printing input
parameters [19]. The mechanical behaviour of 3D-printed materials is influenced by the
diameter of the nozzle used, the material and any filler, the orientation of the layers and
their thicknesses, the feed rate of the nozzle, the angle and width of the grid, and, last
but not least, the density of the filler [20–25]. It is important to describe the mechanical
properties not only by static tests but also by tests simulating cyclic loading, which are
more relevant for practical applications. This is due to the deformation and possible de-
struction of the materials subjected to cyclic loading, which is undesirable. If the cyclic
stress exceeds the elastic limit, plastic deformation accumulates, which then causes the
material to gradually fail [10,18,25–30].
In this article, the suitability of bamboo-filled PLA filament for 3D printing appli-
cations is investigated with emphasis on fatigue behaviour. The mechanical properties
are investigated based on the tensile test results and fatigue behaviours in cyclic tensile
tests on printed test samples using FDM 3D printing technology of bamboo-filled PLA
material. The aim of the research is to determine the basic mechanical characteristics and
fatigue behaviours under different intensities of progressive cyclic loading for tensile tests
for samples with different infill densities and nozzle diameters of 0.4 mm and 0.6 mm.
The effects were studied on 3D printing parameters, including a confirmation whether
the incorporation of a bamboo-based filler into PLA provides the expected mechanical
benefits from a functional point of view, i.e., fatigue loading. The use of bio-based filler
added to the filament is substantial, especially in terms of aesthetic significance visible on
the surface [31].

2. Materials and Methods


2.1. Additive Production of Test Samples
The filament used for the research was a PLA-based filament from (AzureFilm, Sežana,
Slovenia) and a bamboo filler. The material composition was 40% bamboo particles and
60% PLA polymer. The particle size was a trade secret of the company supplying this
filament. The approximate size of the natural filler was determined by optical analysis using
(Gwyddion, Jihlava, Czech Republic) and scanning electron microscopy (SEM) images. The
approximate size of the natural bamboo filler in the PLA matrix was 22.2 ± 8.7 µm. SEM
images before loading were evaluated.
Polymers 2022, 14, 4930 3 of 16

Polymers 2022, 14, x A 3D printer using the Prusa i3 MK3S FDM technology (Prusa Research, a.s., Prague,
4 of 17
Czech Republic) was used to produce the type 1B tensile test samples in accordance with
ČSN EN ISO 527-2 (Figure 1).

Figure 1. Test sample type 1B according to the ČSN EN ISO 527-2 created by 3D printing from
Figure 1. Test sample type 1B according to the ČSN EN ISO 527-2 created by 3D printing from PLA-
PLA-based filament and bamboo filler.
based filament and bamboo filler.
The 3D printer used supports a filament with a diameter of 1.75 mm, so this variant
with the mentioned diameter and a spool weight of 750 g was chosen. The print files
of the test samples were created in the open-source program (PrusaSlicer 2.5.0, Prague,
Czech Republic). Two types of nozzles with diameters of 0.4 mm and 0.6 mm were selected
for printing the samples. The nozzle diameter mainly affected the resulting mechanical
properties of the 3D-printed sample and the overall printing time. Six basic series of
samples were selected, as given in Table 1. The layer height was chosen to be 0.2 mm for all
samples. Another common element was the type of filler chosen. It was a straight filler
applied at an angle of 45◦ to the X-axis. The printing temperature was chosen to be 215 ◦ C
with a heated pad temperature of 60 ◦ C. Printing was carried out in a closed environment
of the print chamber, which reduces the risk of external influences affecting the print.
As a result, the internal temperature in the print chamber was in the range of 22–26 ◦ C
during printing. To compare the weights, a series of pure PLA filament samples with a
straight fill density of 100% was created with the same printing parameters as the previous
samples. The printing of the test samples was relatively easy and smooth, and despite
initial concerns, there was no jamming of the composite filament in the nozzle. Occasional
stringing occurred, which can be observed in Figure 2. The occurrence of this phenomenon
can be2.reduced
Figure by kit
3D printing properly adjusting0.4
for PLA/B/nozzle themm/60%
retractions in the (PrusaSlicer, Prague, Czech
test samples.
Republic). The principle of retraction is to pull the filament back as the print head moves
between
2.2. objects,
Mechanical but this can affect the print quality. In this case, the filaments created did
Tests
not cause problems and could be easily removed.
The mechanical properties of type 1B tensile test samples were investigated based on
the results of tensile
Table 1. Parameters for tests and fatigue
the printing samples. behaviours at different intensities of progressive
cyclic tensile loading for tensile test samples with different infill densities and nozzle di-
ameters Printing
of 0.4 and 0.6 mm. For practical
Variant
Nozzle application,
ø Infillit is the testing
Printingof cyclicAmount
loadingof that
is essential, as it provides results that (mm)are usableDensity (%)
in practice Time (min) to static
compared Filament (m)In
tests.
general, it is known
PLA/B/nozzle that cyclic
0.4 mm/60% * tests simulate
0.4 the gradual
60 changes in35 mechanical properties
2.87
duePLA/B/nozzle 0.4 mm/80%
to different loading intensities. 0.4 80 38 3.15
PLA/B/nozzle
Mechanical 0.4 mm/100%
tests were performed 0.4at a laboratory
100 temperature39of 22 ± 2 °C on3.43 auniver-
sal test machine (LABTest 5.50 ST with an AST KAF6050 kN measuring
PLA/B/nozzle 0.6 mm/60% 0.6 25 2.97
unit (LABOR-TECH
PLA/B/nozzle 0.6 mm/80% 0.6 80 27 3.23
s.r.o., Opava, Czech
PLA/B/nozzle Republic) with
0.6 mm/100% 0.6Test and Motion
100 software27 (LABORTECH 3.48 s.r.o.,
Opava, Czech Republic), enabling cyclic tests with controlled voltage
* PLA—polylactic acid/B—bamboo filler/nozzle 0.4 mm—used nozzle/60%—infill density. modes and load
speeds. The evaluation software Test and Motion allows editing and specific settings of
the fatigue test, and it also records stress and strain data during the experiment. The find-
ings from the static tensile test results were used to determine the baseline parameters
Polymers 2022, 14, 4930 4 of 16
Figure 1. Test sample type 1B according to the ČSN EN ISO 527-2 created by 3D printing from PLA-
based filament and bamboo filler.

Figure 2.
Figure 2. 3D
3D printing
printing kit
kit for
for PLA/B/nozzle
PLA/B/nozzle 0.4
0.4mm/60%
mm/60%test
testsamples.
samples.

Two outer perimeters


2.2. Mechanical Tests with 100% fill were chosen for all printed samples. For the solids
with less
The mechanical propertiesfull
than 100% fill density, fill was
of type 1B chosen for the
tensile test four lower
samples and four upper
were investigated layers
based on
(4 × 0.2 mm = 0.8 mm) due to the sample confinement.
the results of tensile tests and fatigue behaviours at different intensities of progressive
cyclic tensile loading for tensile test samples with different infill densities and nozzle di-
2.2. Mechanical Tests
ameters of 0.4 and 0.6 mm. For practical application, it is the testing of cyclic loading that
The mechanical
is essential, properties
as it provides of type
results that 1B
aretensile
usabletest samples compared
in practice were investigated
to staticbased
tests.on
In
the results of tensile tests and fatigue behaviours at different intensities of progressive
general, it is known that cyclic tests simulate the gradual changes in mechanical properties cyclic
tensile
due to loading
differentforloading
tensile intensities.
test samples with different infill densities and nozzle diameters of
0.4 and 0.6 mm. For practical
Mechanical tests were performedapplication,
at aitlaboratory
is the testing of cyclic loading
temperature of 22 ± 2that is essential,
°C on auniver-
as it provides results that are usable in practice compared to static tests. In general, it
sal test machine (LABTest 5.50 ST with an AST KAF 50 kN measuring unit (LABOR-TECH
is known that cyclic tests simulate the gradual changes in mechanical properties due to
s.r.o., Opava, Czech Republic) with Test and Motion software (LABORTECH s.r.o.,
different loading intensities.
Opava, Czech Republic), enabling cyclic tests with controlled voltage modes and load
Mechanical tests were performed at a laboratory temperature of 22 ± 2 ◦ C on auniver-
speeds. The evaluation software Test and Motion allows editing and specific settings of
sal test machine (LABTest 5.50 ST with an AST KAF 50 kN measuring unit (LABOR-TECH
the fatigue test, and it also records stress and strain data during the experiment. The find-
s.r.o., Opava, Czech Republic) with Test and Motion software (LABORTECH s.r.o., Opava,
ings from the static tensile test results were used to determine the baseline parameters
Czech Republic), enabling cyclic tests with controlled voltage modes and load speeds. The
evaluation software Test and Motion allows editing and specific settings of the fatigue test,
and it also records stress and strain data during the experiment. The findings from the
static tensile test results were used to determine the baseline parameters under low cyclic
loading of the bamboo-filled PLA-based material, i.e., maximum load and strain.
The low-cycle loading of the test samples consisted of successive loads of 250 cycles
at loads from 5 to 20, 30, 40, and 50% based on the limits found in the static tensile test
(reference values used for the cyclic tests).
The cyclic loading, i.e., 1000 cycles, was divided into four consecutive cycles; the first
was at 5–20% loading with 250× repetition, followed by a second loading cycle with 5–30%
loading with 250× repetition, followed by a third cycle with loading in the interval of 5–40%
with 250× repetition, and finally a set of cyclic loading of 5–50% with 250× repetition. A
detailed description of the load variation for each cycle based on the results of the limit
values found in the static tensile test can be seen in Table 2.
Polymers 2022, 14, 4930 5 of 16

Table 2. Low-cycle test setup parameters for each experiment variant tested.

Printing Variant Description of the Low-Cycle Stresses for the Individual Printed Test Variants
The low-cycle loading was from 5% (45 N) to 20% (152 N) with a repetition of 250 cycles, followed by a
PLA/B/nozzle 0.4 mm/60% second loading cycle from 5% to 30% (228 N) 250 cycles, followed by a third loading cycle from 5% to
40% (304 N) 250 cycles, and a final loading cycle from 5% to 50% (379 N) 250 cycles.
The low-cycle loading was from 5% (45 N) to 20% (181 N) with a repetition of 250 cycles, followed by a
PLA/B/nozzle 0.4 mm/80% second loading cycle from 5% to 30% (271 N) 250 cycles, followed by a third loading cycle from 5% to
40% (362 N) 250 cycles, and a final loading cycle from 5% to 50% (452 N) 250 cycles.
The low-cycle loading was from 5% (50 N) to 20% (199 N) with a repetition of 250 cycles, followed by a
PLA/B/nozzle 0.4 mm/100% second loading cycle from 5% to 30% (299 N) 250 cycles, followed by a third loading cycle from 5% to
40% (399 N) 250 cycles, and a final loading cycle from 5% to 50% (499 N) 250 cycles.
The low-cycle loading was from 5% (44 N) to 20% (178 N) with a repetition of 250 cycles, followed by a
PLA/B/nozzle 0.6 mm/60% second loading cycle from 5% to 30% (267 N) 250 cycles, followed by a third loading cycle from 5% to
40% (356 N) 250 cycles, and a final loading cycle from 5% to 50% (445 N) 250 cycles.
The low-cycle loading was from 5% (52 N) to 20% (206 N) with a repetition of 250 cycles, followed by a
PLA/B/nozzle 0.6 mm/80% second loading cycle from 5% to 30% (309 N) 250 cycles, followed by a third loading cycle from 5% to
40% (412 N) 250 cycles, and a final loading cycle from 5% to 50% (515 N) 250 cycles.
The low-cycle loading was from 5% (58 N) to 20% (232 N) with a repetition of 250 cycles, followed by a
PLA/B/nozzle 0.6 mm/100% second loading cycle from 5% to 30% (348 N) 250 cycles, followed by a third loading cycle from 5% to
40% (464 N) 250 cycles, and a final loading cycle from 5% to 50% (580 N) 250 cycles.

The tests were performed at a loading rate of 10 mm × min−1 between 5%, i.e., the
lower limit from the established tensile static test reference value, and the upper limits of
20%, 30%, 40%, and 50%. The relaxation between every set of 1000 cycles or the endurance
at the lower and upper limits was set at 0.5 s. If no destruction of the tested test sample
occurred during cyclic tests, then loading was followed until the limit state was reached,
i.e., destruction with the same loading rate of 10 mm × min−1 . The cyclic fatigue tests
resulted in the ultimate strength, deformation, and deformation differences between the
Polymers 2022, 14, x 6 of 17
1st and 1000th cycles to determine the viscoelastic behaviours. Figure 3 shows the cyclic
testing progress of the test samples.

Figure 3. Low-cycle fatigue


Figure 3. behaviour of 3D-printed
Low-cycle fatigue PLA reinforced
behaviour of 3D-printed withwith
PLA reinforced natural
natural bamboo reinforcement.
bamboo reinforce-
ment.
2.3. Structural Characterization
2.3. Structural Characterization
The test samples that were
The test destroyed
samples duringduring
that were destroyed mechanical tests
mechanical tests were subjected
were subjected to im-to image
analysis of the surfaces using
age analysis a MIRA
of the 3 TESCAN
surfaces using GMX SE
a MIRA 3 TESCAN GMX scanning
SE scanningelectron microscope
electron micro-
scope (Tescan Brno s.r.o., Brno, Czech Republic) with an accelerating voltage of 10 kV and
an Oxford SE detector (Tescan Brno s.r.o, Brno, Czech Republic), while the tested surface
of the test samples was gold-plated for SEM analysis with Quorum Q150R ES—Sputtering
Deposition Rate (Tescan Brno s.r.o., Brno, Czech Republic).

3. Results and Discussion


Polymers 2022, 14, 4930 6 of 16

(Tescan Brno s.r.o., Brno, Czech Republic) with an accelerating voltage of 10 kV and an Oxford
SE detector (Tescan Brno s.r.o., Brno, Czech Republic), while the tested surface of the test
samples was gold-plated for SEM analysis with Quorum Q150R ES—Sputtering Deposition
Rate (Tescan Brno s.r.o., Brno, Czech Republic).

3. Results and Discussion


3.1. Results of Mechanical Tests
The tensile strength of test samples printed by FDM technology from PLA is about
60 MPa [1]. The previously reported research results show that with the addition of different
fillers to PLA, the tensile strength usually decreases. This decrease in tensile strength then
increases with filler concentration [5,25,30,31]. The degree of decrease in tensile strength
depends on the type, concentration, and shape of the filler. Azadi et al. reported that PLA has
a better fatigue life than samples made from other polymers, e.g., ABS [18].
Another significant factor affecting the mechanical behaviour of 3D-printed products is
infill density. The results of the basic mechanical characteristics of the tested bamboo-filled
PLA filament-based materials, i.e., tensile strength (Figure 4) and elongation at break (Figure 5),
present different behaviours in static testing at different infill densities and using two different
nozzles of 0.4 and 0.6 mm. The static tensile test results confirmed the significant effect of
infill densities of 60%, 80%, and 100% on the tensile strength of the tested samples. Using a
0.4 mm nozzle, the static tensile strength at 100% infill density (PLA/B/nozzle 0.4 mm/100%)
was 24.6 ± 0.9 MPa. The results (Figure 4) show an almost linear trend of decrease in tensile
strength with decreasing infill density. For 80% infill density there was a 10% decrease in
tensile strength, and for 60% infill density there was a decrease of up to 24%. A similar trend
was found when using a nozzle diameter of 0.6 mm. The static tensile strength at 100% infill
density (PLA/B/nozzle 0.6 mm/100%) was 27.9 ± 0.3 MPa. There was a 15% decrease in
tensile strength at 80% infill density using a 0.6 mm nozzle diameter and up to a 25% decrease
with 60% infill density. The arithmetic mean of the tensile strength results also shows that there
was a 6–12% increase in tensile strength when using a larger nozzle diameter, i.e., 0.6 mm.
It is evident from the research on adding fillers to PLA filaments that apart from a
reduction in strength, there is usually also an increase in elongation at break [25,30,31]. The
elongation at break for the tested samples printed by FDM from PLA was about 3.5% [1].
Figure 5 presents the elongation at break for static testing at different infill densities
and using two different nozzles of 0.4 and 0.6 mm. From the results, it is clear that there is
no significant change due to different nozzle diameters and infill densities of 60, 80, and
100% when considering the variance of the results plotted in Figure 5. It is also evident
from the results of the arithmetic mean of elongation at break that there is no significant
change when using a larger nozzle diameter. When a nozzle diameter of 0.4 mm was used,
the elongation at break was 4.16 ± 0.26% at 100% infill density for the sample labelled
PLA/B/nozzle 0.4 mm/100%. At 80% and 60% infill densities, there was an identical 13%
increase in elongation at break. Using a 0.6 mm diameter nozzle, elongation at break was
3.97 ± 0.87% at 100% infill density for the sample labelled PLA/B/nozzle 0.6 mm/100. At
80% infill density, using a 0.6 mm diameter nozzle, there was a 4% decrease in elongation
at break, but at 60% infill density, there was a 4.8% increase in elongation at break.
density there was a 10% decrease in tensile strength, and for 60% infill density there was
a decrease of up to 24%. A similar trend was found when using a nozzle diameter of 0.6
mm. The static tensile strength at 100% infill density (PLA/B/nozzle 0.6 mm/100%) was
27.9 ± 0.3 MPa. There was a 15% decrease in tensile strength at 80% infill density using a
0.6 mm nozzle diameter and up to a 25% decrease with 60% infill density. The arithmetic
Polymers 2022, 14, 4930 7 of 16
mean of the tensile strength results also shows that there was a 6–12% increase in tensile
strength when using a larger nozzle diameter, i.e., 0.6 mm.

Polymers 2022, 14, x 8 of 17


Figure 4.
Figure Influence of
4. Influence of infill
infill density
density on
on tensile
tensile strength
strength in
in static
static testing
testing using
using FDM
FDM 3D
3D printing
printing with
with
different nozzle diameters.

It is evident from the research on adding fillers to PLA filaments that apart from a
reduction in strength, there is usually also an increase in elongation at break [25,30,31].
The elongation at break for the tested samples printed by FDM from PLA was about 3.5%
[1].
Figure 5 presents the elongation at break for static testing at different infill densities
and using two different nozzles of 0.4 and 0.6 mm. From the results, it is clear that there
is no significant change due to different nozzle diameters and infill densities of 60, 80, and
100% when considering the variance of the results plotted in Figure 5. It is also evident
from the results of the arithmetic mean of elongation at break that there is no significant
change when using a larger nozzle diameter. When a nozzle diameter of 0.4 mm was used,
the elongation at break was 4.16 ± 0.26% at 100% infill density for the sample labelled
PLA/B/nozzle 0.4 mm/100%. At 80% and 60% infill densities, there was an identical 13%
increase in elongation at break. Using a 0.6 mm diameter nozzle, elongation at break was
3.97 ± 0.87% at 100% infill density for the sample labelled PLA/B/nozzle 0.6 mm/100. At
80% infill density, using a 0.6 mm diameter nozzle, there was a 4% decrease in elongation
at break, but at 60% infill density, there was a 4.8% increase in elongation at break.

Figure 5.
Figure Effect of
5. Effect of infill
infill density
density on
on elongation
elongation at
at break
break in
in static
static testing
testing using
using FDM
FDM 3D
3D printing
printing with
with
different nozzle diameters.
different nozzle diameters.

The results
The results ofofthe
thetensile
tensilestrength
strength after thethe
after low-cycle test test
low-cycle can be
canseen
be in Figure
seen 6, which
in Figure 6,
which consisted of 1000 cycles divided into four consecutive load intervals from 540,
consisted of 1000 cycles divided into four consecutive load intervals from 5 to 20, 30, toand
20,
50%40,
30, based on the
and 50% limiton
based values found
the limit in thefound
values staticin
tensile test divided
the static intodivided
tensile test 250 consecutive
into 250
cycles. The research results showed minimal effects on the low-cycle
consecutive cycles. The research results showed minimal effects on the low-cycle tensile properties
tensile
according to the conditions defined in Table 2 using additive manufacturing
properties according to the conditions defined in Table 2 using additive manufacturing by 3D printing
using the FDM method from PLA-based filament and bamboo filler. All test samples
by 3D printing using the FDM method from PLA-based filament and bamboo filler. All
withstood successive cyclic loading at different loading force intensities (Table 3). The
test samples withstood successive cyclic loading at different loading force intensities (Ta-
results are consistent with previous findings dealing with fatigue life and showed a positive
ble 3). The results are consistent with previous findings dealing with fatigue life and
showed a positive or neutral effect of the filler. For all experimental variations, low-cycle
testing resulted in an increase in tensile strength over the static tensile test values ranging
from 0.2 to 9.4% [10,28]. This trend is demonstrated by Senatov et al. in their research, i.e.,
there is a cyclic strengthening of the material [29]. The higher strengthening after low-
Polymers 2022, 14, 4930 8 of 16

or neutral effect of the filler. For all experimental variations, low-cycle testing resulted
in an increase in tensile strength over the static tensile test values ranging from 0.2 to
9.4% [10,28]. This trend is demonstrated by Senatov et al. in their research, i.e., there is a
cyclic strengthening of the material [29]. The higher strengthening after low-cycle testing
relative to the static test values occurred for 3D printing with a nozzle of 0.4 mm, i.e., in
the interval from 1.8 to 9.4%. Another finding is that the increase of 2.9 to 9.4% in tensile
strength after low-cycle testing occurred for materials with an infill density of 100%. As the
infill density decreased, the final tensile strength after low-cycle test decreased. From the
Polymers 2022, 14, x 9 of 17
point of view of practical application, the finding that low-cycle fatigue does not reduce the
tensile strength up to 50% of the load is significant.

Effectof
Figure6.6.Effect
Figure ofinfill
infilldensity
densityon
ontensile
tensilestrength
strengthin
inlow-cycle
low-cycletesting
testingusing
usingFDM
FDM3D
3Dprinting
printingwith
with
different nozzle diameters.
different nozzle diameters.
Table 3. Results of deformation of the test samples between the first and last test cycles in the
From the results of the elongation at break after low-cycle testing observed in Figure
low-cycle test.
7, it is clear that there is no significant change due to the change of the nozzle diameter
and the infill densities of 60%, 80%, and 100% considering the varianceStrain
Relative of theDifference
results. The
Relative Deformation
Printing Variant Numberresults of theTests
of Finished arithmeticDeformation
mean of the1st elongation at break after low-cyclebetweentest show 1staand
similar
1000th Cycle (%)
trend to that found in the Cycle
static (%)
tensile test (Figure 5). When a larger nozzle 1000thdiameter,
Cycle (%) i.e.,
PLA/B/nozzle 0.4 0.6 mm, was used, the elongation at break after low-cycle testing was reduced by 5 to 16%.
1000 0.069 ± 0.005 0.174 ± 0.005 0.11 ± 0.01
mm/60% When a nozzle diameter of 0.4 mm was used, the elongation at break after low-cycle test-
PLA/B/nozzle 0.4 ing1000
was 4.46 ± 0.19% at0.072 100% infill density for the±sample
± 0.005 0.165 0.014 labelled PLA/B/nozzle
0.09 ± 0.01 0.4
mm/80% mm/100%. At 80% and 60% infill densities, there was an increase in elongation at break
PLA/B/nozzle 0.4
after low-cycle testing of0.065
1000 3.8 to 4.1%. Using a 0.60.159
± 0.000 mm±diameter
0.016 nozzle, the0.09elongation
± 0.02 at
mm/100%
PLA/B/nozzle 0.6 break after low-cycle testing was 4.15 ± 0.37% for the 100% infill density sample labelled
mm/60%
1000
PLA/B/nozzle 0.6 mm/100. At±80%
0.069 0.005 using±a0.011
infill density 0.171 0.6 mm nozzle, there 0.11 ± 0.01an 11%
was
PLA/B/nozzle 0.6 decrease in elongation at break after the low-cycle testing, and at 60% infill density there
1000 0.072 ± 0.005 0.184 ± 0.005 0.11 ± 0.00
mm/80% was also a 2.0% decrease.
PLA/B/nozzle 0.6 The low-cycle testing resulted
1000 0.078 ± 0.005in an increase in elongation
0.184 ± 0.005 at break compared
0.10 ± 0.01 to the
mm/100%
static tensile test values only for 100% infill density, ranging from 4.6 to 7.1%. For the other
experimental variants having 80 and 60% infill densities, the low-cycle test resulted in a
decrease
Thein elongation
results of theat break
static relative
tensile testtoafter
the values found in
the low-cycle thealso
test static test. These
confirmed theresults
signif-
icant effect
showed of the infill
a decrease densities of
in deformation 60%,low-cycle
under 80%, andtesting.
100% on the tensile
When using astrengths of the
0.4 mm nozzle
tested
the samples. The
deformation was results (Figure
decreased from6)9.1
show an almost
to 9.3%. Whenlinear
usingtrend of decreases
a 0.6 mm nozzle the in defor-
tensile
mation decreased from 2.1 to 3.0%.
Polymers 2022, 14, 4930 9 of 16

strength after low-cycle testing with decreasing infill densities. Using a 0.4 mm diameter
nozzle, the tensile strength after low-cycle testing for 100% infill density (PLA/B/nozzle
0.4 mm/100%) was 27.0 ± 0.5 MPa. At 80% infill density, the tensile strength after the
low-cycle test decreased by 15%, and at 60% infill density, it decreased by up to 29.6%.
A similar trend was found when using a 0.6 mm diameter nozzle. The tensile strength
after low-cycle testing at 100% infill density (PLA/B/nozzle 0.6 mm/100%) was 28.7 ± 0.3
MPa. At 80% infill density there was a decrease in tensile strength after low-cycle testing
by 17.6%, and at 60% infill density, it decreased by up to 26.2%. The results presented by
Essassi et al. also showed the importance of density on fatigue properties [30].
From the results of the elongation at break after low-cycle testing observed in Figure 7,
it is clear that there is no significant change due to the change of the nozzle diameter and
the infill densities of 60%, 80%, and 100% considering the variance of the results. The
results of the arithmetic mean of the elongation at break after low-cycle test show a similar
trend to that found in the static tensile test (Figure 5). When a larger nozzle diameter, i.e.,
0.6 mm, was used, the elongation at break after low-cycle testing was reduced by 5 to
16%. When a nozzle diameter of 0.4 mm was used, the elongation at break after low-cycle
testing was 4.46 ± 0.19% at 100% infill density for the sample labelled PLA/B/nozzle
0.4 mm/100%. At 80% and 60% infill densities, there was an increase in elongation at break
after low-cycle testing of 3.8 to 4.1%. Using a 0.6 mm diameter nozzle, the elongation at
break after low-cycle testing was 4.15 ± 0.37% for the 100% infill density sample labelled
Polymers 2022, 14, x
PLA/B/nozzle 0.6 mm/100. At 80% infill density using a 0.6 mm nozzle, there was 10 anof11%
17
decrease in elongation at break after the low-cycle testing, and at 60% infill density there
was also a 2.0% decrease.

Figure 7.
Figure Effect of
7. Effect of infill
infill density
density on
on elongation
elongation at
at break
break in
in low-cycle testing using
low-cycle testing using FDM
FDM 3D
3D printing
printing
with different nozzle diameters.

The low-cycle
Table 3 shows testing resulted
the difference inin an increasebetween
deformation in elongation at break
the first and the compared
last (i.e., to the
1000)
static tensile test values only for 100% infill density, ranging from 4.6 to 7.1%.
test cycles for each tested variant. This difference between the first and last test cycles For the
other experimental
represents the shift variants having 80loops
of the hysteresis and 60% infilllow-cycle
for the densities,tests
the low-cycle
(Figures 8test
andresulted
9) and
in a decrease in elongation at break relative to the values found in the
represents the viscoelastic behaviour or creep of the tested experimental variants. static test. Exam-
These
ples of the loading curves and low-cycle fatigue test progress are shown in Figure 8mm
results showed a decrease in deformation under low-cycle testing. When using a 0.4 for
nozzle
the 0.4 the
mmdeformation was decreased
diameter nozzle from9 9.1
and in Figure fortothe9.3%. When
0.6 mm using anozzle.
diameter 0.6 mmThe nozzle the
results
deformation decreased from 2.1 to 3.0%.
show the deformation between the first and last test cycles (1000) of the test specimens
Table 3 shows the difference in deformation between the first and the last (i.e., 1000)
during the low-cycle test, which presents the fatigue behaviour of the printed test samples
test cycles for each tested variant. This difference between the first and last test cycles
by FDM 3D printing technology of bamboo-filled PLA material at different infill densities
and using two nozzle diameters. The difference between the 1st and 1000th cycle ranged
from 0.09 to 0.11%. From the results shown in Table 3, there is no clear trend of the differ-
ence between the 1st and 1000th cycle depending on the nozzle diameter and infill density.
It can only be stated that the 0.6 mm nozzle showed similar values of difference between
Polymers 2022, 14, 4930 10 of 16

represents the shift of the hysteresis loops for the low-cycle tests (Figures 8 and 9) and
represents the viscoelastic behaviour or creep of the tested experimental variants. Examples
of the loading curves and low-cycle fatigue test progress are shown in Figure 8 for the
Polymers 2022, 14, x 0.4 mm diameter nozzle and in Figure 9 for the 0.6 mm diameter nozzle. The results11show of 17
the deformation between the first and last test cycles (1000) of the test specimens during
the low-cycle test, which presents the fatigue behaviour of the printed test samples by
FDM
Table 3D printing
3. Results technology of
of deformation of bamboo-filled
the test samplesPLA
betweenmaterial at different
the first infill
and last test densities
cycles and
in the low-
using two nozzle diameters. The difference between the 1st and 1000th cycle ranged from
cycle test.
0.09 to 0.11%. From the results shown in Table 3, there is no clear trend of the difference
between the 1st and 1000thNumber cycle depending on the nozzle diameter Strain
and infill Differ-It
density.
of Relative Relative Defor-
can only be stated that the 0.6 mm nozzle showed similar values of difference ence between
between
Printing Variant Finished Deformation mation 1000th
the 1st and 1000th cycle at different infill densities. However, it is clear from 1st and
the1000th
results
Tests 1st Cycle (%) Cycle (%)
that the low-cycle test showed a creep of the test samples during loading. This Cycle
creep(%)was
manifested by a deformation
PLA/B/nozzle 0.4 characterized by the difference in strain between the first
and last cycle. This trend is also1000evident 0.069
from ±the0.005
shift of0.174 ± 0.005 loops
the hysteresis 0.11shown
± 0.01 in
mm/60%
Figures 8 and 9. After
PLA/B/nozzle 0.4stress relief, the hysteresis curve should tend to return to its original
state. Thismm/80%
would prevent material 1000creep. The
0.072accumulated
± 0.005 0.165
stress±during
0.014 low-cycle
0.09 ± fatigue
0.01
induces deformation that leads to different magnitudes of hysteresis loop displacement,
PLA/B/nozzle 0.4
which can be seen from Figures1000 8 and 9 and0.065
also± from
0.000the results
0.159 ±shown
0.016 in Table0.093,± i.e.,
0.02the
mm/100%
difference in deformation between cycles 1 and 1000.
PLA/B/nozzle 0.6
If this strain exceeded the1000 elastic limit, significant
0.069 ± 0.005 plastic
0.171deformation
± 0.011 would
0.11 occur,
± 0.01
mm/60%
leading to premature destruction of the test samples [1]. However, this did not happen,
PLA/B/nozzle
and all test samples0.6 withstood1000the specified 1000 load cycles.
0.072 ± 0.005 0.184Tao and Xia stated
± 0.005 0.11 ±that
0.00the
mm/80% of loading will have a significant effect on the fatigue behaviours of
rate and progression
PLA/B/nozzle
polymeric materials.0.6Therefore, it is very important to obtain strain and stress data during
1000 0.078 ± 0.005 0.184 ± 0.005 0.10 ± 0.01
mm/100%
the fatigue process [32].

Influenceofofinfill
Figure 8. Influence infilldensity
densityonon low-cycle
low-cycle fatigue
fatigue using
using FDMFDM
3D 3D printing
printing withwith
a 0.4amm
0.4 noz-
mm
zle diameter.
nozzle diameter.
Polymers 2022, 14, x4930 1211ofof 17
16

Figure 9.
Figure Influenceofofinfill
9.Influence infill density
density onon low-cycle
low-cycle fatigue
fatigue using
using FDMFDM 3D printing
3D printing withwith
a 0.6 amm
0.6noz-
mm
zle diameter.
nozzle diameter.

3.2. SEM
3.2. SEM Analysis
Analysis ofof Tested
Tested Materials
Materials
SEM images
SEM images of of the
the fracture
fracture surface
surface using
using 3D 3D printing
printing with
with FDM
FDM of of aa bamboo-filled
bamboo-filled
PLA-based material with infill densities and nozzle diameters
PLA-based material with infill densities and nozzle diameters of 0.4 and 0.6 mm of 0.4 and 0.6 mm cancan bebe
seen in Figures 10–14. These images present the interactions of the bamboo fillers and
seen in Figures 10–14. These images present the interactions of the bamboo fillers and PLA
PLA matrices, the structures and sizes of the layers affected by infill density and nozzle
matrices, the structures and sizes of the layers affected by infill density and nozzle diam-
diameter, and finally the changes caused by low-cycle fatigue. From the fracture surface
eter, and finally the changes caused by low-cycle fatigue. From the fracture surface shown
shown in Figure 10, a good interaction of the bamboo filler with the matrix can be seen
in Figure 10, a good interaction of the bamboo filler with the matrix can be seen when
when using nozzle diameters of 0.4 and 0.6 mm. The results are consistent with the claim
using nozzle diameters of 0.4 and 0.6 mm. The results are consistent with the claim that
that the biofiller has a good interfacial interaction with the PLA matrix [27]. This is the
the biofiller has a good interfacial interaction with the PLA matrix [27]. This is the first
first prerequisite for the successful use of the tested bamboo-filled PLA-based filament.
prerequisite for the successful use of the tested bamboo-filled PLA-based filament. During
During 3D printing by the FDM method, rapid diffusion occurs during high-temperature
3D printing by the FDM method, rapid diffusion occurs during high-temperature pro-
processing, which causes interdiffusion bonding between the biofiller and PLA, which may
cessing, which causes interdiffusion bonding between the biofiller and PLA, which may
result in good contact between the matrix and the filler and could improve the resulting
result in good contact between the matrix and the filler and could improve the resulting
mechanical properties [10].
mechanical
Figuresproperties
11 and 12 [10].
are examples of figures showing the connection of individual layers,
Figures
their arrangement,11 andand12 are examples
their porosity.ofFigures
figures 12 showing
and 13 the
show connection
consistentlyof individual
good bonding lay-
ers, their arrangement, and their porosity. Figures 12 and 13 show
of the individual layers. The assumption of Chen et al. that poor interlayer bonding consistently good bond-
ing of the
occurs whenindividual
using FDM layers. The assumption
technology, leading toofanisotropy
Chen et al.inthat poor interlayer
mechanical properties,bonding
is not
occurs when using FDM technology, leading to anisotropy in mechanical
confirmed [33]. Figures 11 and 12A,B show the porosity that usually occurs in nonideal properties, is
not confirmed [33]. Figures 11 and 12A,B show the porosity that usually
manufacturing during 3D printing and in materials with natural fillers [30,34]. The weight occurs in nonideal
manufacturing
of the test sample during 3D printing
fabricated by 3Dand in materials
printing from PLA withwas
natural fillers
9.95 g. For [30,34].
PLA with Thebamboo-
weight
of the test sample fabricated by 3D printing from PLA was 9.95 g. For PLA
based filler, the weight ranged from 7.40 g to 8.91 g. The weight varied due to infill density. with bamboo-
based filler,anthe
There was weightinranged
increase weightfrom with7.40 g to 8.91
increasing g. The
infill weight
density varied
value. due to
At 100% infill
infill den-
density,
sity. There was
the weight wasan increase
found to bein8.87
weight withaincreasing
g using infill density
0.4 mm diameter nozzle value. At 100%
and 8.91 g usinginfilla
density, the weight
0.6 mm diameter was found to be 8.87 g using a 0.4 mm diameter nozzle and 8.91 g
nozzle.
using a 0.6 mm diameter nozzle.
Polymers 2022, 14, 4930
x 1312of
of 17
16
Polymers 2022, 14, x 13 of 17
Polymers 2022, 14, x 13 of 17

Figure 10.
10. SEM images—fractographic
images—fractographic analysis
analysis (MAG
(MAG 2.00
2.00 kx):
kx): (A):
(A): nozzle
nozzle 0.4
0.4 mm,
mm, infill density
Figure
Figure 10. SEM
SEM images—fractographic
images—fractographic analysis (MAG 2.00
2.00 kx): (A): nozzle 0.4 mm, infill density
60%, static test; (B): nozzle 0.4 mm, infillanalysis
density(MAG
80%, static kx):
test; (A):
(C): nozzle
nozzle 0.4
0.6 mm,
mm, infill density
infill density
60%,
60%, static
static test;
test; (B):
(B): nozzle
nozzle 0.4
0.4 mm,
mm, infill
infill density
density 80%,
80%, static
static test;
test; (C):
(C): nozzle
nozzle 0.6
0.6 mm,
mm, infill
infill density
density
100%, low cyclic fatigue.
100%,
100%, low
100%, low cyclic
low cyclic fatigue.
cyclic fatigue.
fatigue.

Figure 11. SEM images—fractographic analysis—nozzle 0.6 mm, infill density 60%, static tensile
Figure 11. SEM
SEM images—fractographic
images—fractographic analysis—nozzle
analysis—nozzle 0.6
0.6 mm,
mm, infill
infill density 60%, static tensile
Figure 11.
test—connection images—fractographic analysis—nozzle
of individual layers, arrangements, 0.6
and theirmm, infill
porosities: density
(A): MAG60%, (B):tensile
static
150×; MAG
test—connection
test—connection of individual
individual layers, arrangements,
layers, arrangements, and
arrangements,and their
andtheir porosities:
theirporosities: (A):
porosities:(A): MAG
(A):MAG 150×;
MAG150
150×; (B): MAG
500×; of individual
(C): MAG 1.00 kx. layers, ×; (B): MAG
500×;
500×; (C): MAG 1.00 kx.
500×;(C):
(C):MAG
MAG1.00
1.00kx.
kx.

Figure 12. SEM images—fractographic analysis—nozzle 0.4 mm, infill density 60%, low cyclic fa-
Figure
Figure 12. SEM images—fractographic analysis—nozzle 0.4 mm, infill density 60%, low cyclic fa-
Figure 12.
12. SEM
SEMimages—fractographic
tigue—connection
tigue—connection
of individual layers,analysis—nozzle
images—fractographic
of individual layers,
arrangements,
analysis—nozzle
arrangements,
0.40.4
and
and
mm,
their
mm,
their
infill density
porosities:
infill 60%,
(A):
density
porosities: (A):
lowlow
MAG
60%,
MAG
cyclic
150×;
150×;
fa-
(B):
cyclic
(B):
tigue—connection
MAG of individual
500×; (C): MAGof1.00
fatigue—connection kx. layers,
individual arrangements, and their porosities: (A): MAG 150×;
layers, arrangements, and their porosities: (A): MAG 150×; (B): (B):
MAG
MAG 500×;
500×; (C):
(C): MAG
MAG 1.00
1.00 kx.
kx.
MAG 500×; (C): MAG 1.00 kx.
Figures 13 and 14 show a comparison of 3D printing with nozzle diameters of 0.4 and
Figures
Figures 13
13 and
and 14
14 show
show aa comparison
comparison of 3D printing
printingofwith nozzle
nozzle diameters of
of 0.4 and
0.6 mm. Figures 13 and 14A–C present theof
SEM3D results with diameters
the fractographic 0.4after
analysis and
0.6
0.6 mm.
mm. Figures
Figures 13
13 and
and 14A–C
14A–C present
present the
the SEM
SEM results
results of
of the
the fractographic
fractographic analysis
analysis after
after
tensile testing. Figures 13 and 14D–F present the results of SEM fractographic analysis
tensile
tensile testing.
testing. Figures
Figures 13
13 and
and 14D–F
14D–F present
present the
the results
results of
of SEM
SEM fractographic
fractographic analysis
analysis
PLA/B/nozzle 0.4 mm/60% 0.22 ± 0.02 0.23 ± 0.02
PLA/B/nozzle 0.4 mm/80% 0.20 ± 0.02 0.22 ± 0.32
PLA/B/nozzle 0.4 mm/100% 0.18 ± 0.03 0.18 ± 0.02
PLA/B/nozzle 0.6 mm/60% 0.21 ± 0.02 0.21 ± 0.03
Polymers 2022, 14, 4930 PLA/B/nozzle 0.6 mm/80% 0.20 ± 0.02 0.20 ± 0.01 13 of 16

PLA/B/nozzle 0.6 mm/100% 0.18 ± 0.02 0.19 ± 0.02

Figure 13.
Figure 13. SEM
SEM images—fractographic
images—fractographic analysis—nozzle
analysis—nozzle 0.40.4 mm
mm (MAG
(MAG2.50 2.50×):
×): (A): Infill density
(A): Infill density
100%, static tensile test; (B): Infill density 80%, static tensile test; (C): Infill density 60%, static tensile
100%, static tensile test; (B): Infill density 80%, static tensile test; (C): Infill density 60%, static tensile
Polymers 2022, 14, x test; (D): Infill density 100%, low cyclic fatigue; (E): Infill density 80%, low cyclic fatigue; (F): 15 of 17
Infill
test; (D): Infill density 100%, low cyclic fatigue; (E): Infill density 80%, low cyclic fatigue; (F): Infill
density 60%, low cyclic fatigue.
density 60%, low cyclic fatigue.

14. SEM
Figure 14. SEM images—fractographic
images—fractographic analysis—nozzle
analysis—nozzle 0.6 0.6 mm
mm (MAG
(MAG 2.50 ×): (A): Infill density
2.50×): density
100%, static
100%, static tensile
tensile test;
test; (B):
(B): Infill
Infill density
density 80%,
80%, static
static tensile
tensile test;
test; (C): Infill
Infill density
density 60%, static tensile
test; (D):
test; (D): Infill
Infill density
density 100%,
100%, low
low cyclic
cyclic fatigue;
fatigue; (E):
(E): Infill
Infill density
density 80%,
80%, low
low cyclic
cyclic fatigue;
fatigue; (F):
(F): Infill
Infill
density 60%, low cyclic fatigue.
density 60%, low cyclic fatigue.

4. Conclusions
In this article, results focused on 3D printing using PLA-based filament with natural
bamboo filler are reported. The development of materials with natural filler is very im-
portant in the field of additive technologies. Bamboo-based natural filler contributes sig-
Polymers 2022, 14, 4930 14 of 16

Figures 13 and 14 show a comparison of 3D printing with nozzle diameters of 0.4 and
0.6 mm. Figures 13 and 14A–C present the SEM results of the fractographic analysis after
tensile testing. Figures 13 and 14D–F present the results of SEM fractographic analysis after
low-cycle fatigue. From Figures 13 and 14, the effect of infill densities of 100%, 80%, and
60% can be compared. In their research, Jerez-Mesa et al. found that the layer height had a
significant effect on the fatigue life in PLA [35]. For this reason, the fracture surface seen in
the SEM images was subjected to image analysis using Gwyddion software, the results of
which can be observed in Table 4.

Table 4. Results of layer height measurements determined from SEM image analysis.

Layer Height—Static Layer Height- Low-Cycle


Printing Variant
Test (mm) Fatigue (mm)
PLA/B/nozzle 0.4 mm/60% 0.22 ± 0.02 0.23 ± 0.02
PLA/B/nozzle 0.4 mm/80% 0.20 ± 0.02 0.22 ± 0.32
PLA/B/nozzle 0.4 mm/100% 0.18 ± 0.03 0.18 ± 0.02
PLA/B/nozzle 0.6 mm/60% 0.21 ± 0.02 0.21 ± 0.03
PLA/B/nozzle 0.6 mm/80% 0.20 ± 0.02 0.20 ± 0.01
PLA/B/nozzle 0.6 mm/100% 0.18 ± 0.02 0.19 ± 0.02

Jerez-Mesa et al. reported that with increasing height of the PLA layer of the test
samples, better results were found in the number of cycles for resisting failure [35]. This
was not demonstrated in the research. The test samples withstood the entire low-cycle
fatigue test cycle. From the results shown in Table 4, the trend of the difference in the layer
found from SEM image fractographic analysis during static and low-cycle fatigue testing
is not clear. However, the effect of infill density is evident from the results. There was a
decrease in layer height with an increasing infill density value (Table 4).
There was no reduction in sample height, pore collapse, or delamination due to cyclic
loading, as indicated in the research of Senatov et al. [29].

4. Conclusions
In this article, results focused on 3D printing using PLA-based filament with natu-
ral bamboo filler are reported. The development of materials with natural filler is very
important in the field of additive technologies. Bamboo-based natural filler contributes
significantly to the improvement of aesthetic properties. The results demonstrated the
suitability of bamboo-filled PLA filament for 3D printing applications, with emphasis
on fatigue behaviour under different intensities of successive load cycles. The obtained
results of controlled loading during cyclic tests demonstrated the functionality of the tested
material at the evaluated manufacturing parameters affecting the durability of potential
3D printed products made from PLA-based filament filled with bamboo filler subjected
to low-cycle fatigue. Cyclic stresses represent a common cause of failure of their adhesive
bonds, thereby reducing their service life.
Based on the results presented in this article, the following conclusions can be drawn:
• The static tensile test results confirmed the significant effects of infill densities of 60%,
80%, and 100% on the tensile strength of the tested samples and of 0.4 and 0.6 mm
printing nozzles on the tensile strength. In particular, the research results show the
significant effect of infill density on the fatigue properties of the tested materials.
• For all tested variants of the experiment, the low-cycle test resulted in an increase in
tensile strength compared to the values from the static test, up to about 10%. Thus,
a so-called cyclic strengthening of the material occurred. The viscoelastic behaviour
(creep) of the material during low-cycle fatigue was also evident. Infill density again
proved to have a significant influence. At 100% infill density, the highest fatigue tensile
strength was achieved.
Polymers 2022, 14, 4930 15 of 16

• SEM analysis of the fracture surface confirmed good interaction between the PLA
matrix and the bamboo-based filler using the tested print nozzle diameters of both 0.4
and 0.6 mm.
The results of SEM analysis did not show any reductions in the layer height of the
samples, pore collapse, or delamination due to low cyclic loading. However, they showed
the effect of infill density on the layer height of the samples.

Author Contributions: Conceptualization, M.M., V.Š., P.J. and R.K.M.; methodology, V.Š. and M.M.;
software, P.J., M.M. and J.N.; testing of mechanical properties, V.Š., P.J., M.M., R.K.M. and M.H.;
data analysis, M.M., V.Š., R.K.M. and P.J.; writing—original draft preparation, M.M., P.J. and R.K.M.;
writing—review and editing, M.M., R.K.M. and P.J.; resources, V.Š., M.M., R.K.M., P.J. and M.H.;
performed the SEM analysis, M.M. and M.H.; supervision, M.M. and R.K.M. All authors have read
and agreed to the published version of the manuscript.
Funding: Supported by the internal grant agency of the Faculty of Engineering, Czech University
of Life Sciences Prague grants no. 2022:31140/1312/3105: “Research on composite materials with
polymer matrix and natural filler in the field of additive technology” and by the OP VVV Project De-
velopment of new nano and micro coatings on the surface of selected metallic materials-NANOTECH
ITI II., Reg. No CZ.02.1.01/0.0/0.0/18_069/0010045.
Institutional Review Board Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Muller, M.; Sleger, V.; Kolar, V.; Hromasova, M.; Pis, D.; Mishra, R.K. Low-Cycle Fatigue Behavior of 3D-Printed PLA Reinforced
with Natural Filler. Polymers 2022, 14, 1301. [CrossRef] [PubMed]
2. Ayrilmis, N.; Buyuksari, U.; Dundar, T. Waste Pine Cones as a Source of Reinforcing Fillers for Thermoplastic Composites. J. Appl.
Polym. Sci. 2010, 117, 2324–2330. [CrossRef]
3. Ruggiero, A.; Valasek, P.; Mueller, M. Exploitation of Waste Date Seeds Phoenix Dactylifera in Form of Polymeric Particle
Biocomposite: Investigation on Adhesion, Cohesion. Compos. Part B Eng. 2016, 104, 9–16. [CrossRef]
4. Pakkanen, J.; Manfredi, D.; Minetola, P.; Iuliano, L. About the use of recycled or biodegradable filaments for sustainability of
3D printing. In Sustainable Design and Manufacturing 2017; Campana, G., Howlett, R.J., Setchi, R., Cimatti, B., Eds.; Springer
International Publishing: Cham, Germany, 2017; pp. 776–785.
5. Antonio Travieso-Rodriguez, J.; Zandi, M.D.; Jerez-Mesa, R.; Lluma-Fuentes, J. Fatigue Behavior of PLA-Wood Composite
Manufactured by Fused Filament Fabrication. J. Mater. Res. Technol. 2020, 9, 8507–8516. [CrossRef]
6. Svatik, J.; Lepcio, P.; Ondreas, F.; Zarybnicka, K.; Zboncak, M.; Mencik, P.; Jancar, J. PLA Toughening via Bamboo-Inspired 3D
Printed Structural Design. Polym. Test. 2021, 104, 107405. [CrossRef]
7. Zhao, D.X.; Cai, X.; Shou, G.Z.; Gu, Y.Q.; Wang, P.X. Study on the Preparation of Bamboo Plastic Composite Intend for Additive
Manufacturing. Key Eng. Mater. 2016, 667, 250–258. [CrossRef]
8. Correa, D.; Papadopoulou, A.; Guberan, C.; Jhaveri, N.; Reichert, S.; Menges, A.; Tibbits, S. 3D-Printed Wood: Programming
Hygroscopic Material Transformations. 3D Print. Addit. Manuf. 2015, 2, 106–116. [CrossRef]
9. Ayrilmis, N. Effect of Layer Thickness on Surface Properties of 3D Printed Materials Produced from Wood Flour/PLA Filament.
Polym. Test. 2018, 71, 163–166. [CrossRef]
10. Shahar, F.S.; Sultan, M.T.H.; Safri, S.N.A.; Jawaid, M.; Abu Talib, A.R.; Basri, A.A.; Shah, A.U.M. Fatigue and Impact Properties of
3D Printed PLA Reinforced with Kenaf Particles. J. Mater. Res. Technol. 2022, 16, 461–470. [CrossRef]
11. Yao, T.; Zhang, K.; Deng, Z.; Ye, J. A Novel Generalized Stress Invariant-Based Strength Model for Inter-Layer Failure of FFF 3D
Printing PLA Material. Mater. Des. 2020, 193, 108799. [CrossRef]
12. Le Duigou, A.; Correa, D.; Ueda, M.; Matsuzaki, R.; Castro, M. A Review of 3D and 4D Printing of Natural Fibre Biocomposites.
Mater. Des. 2020, 194, 108911. [CrossRef]
13. Huber, T.; Müssig, J. Fibre Matrix Adhesion of Natural Fibres Cotton, Flax and Hemp in Polymeric Matrices Analyzed with the
Single Fibre Fragmentation Test. Compos. Interfaces 2008, 15, 335–349. [CrossRef]
14. Arockiam, A.J.; Subramanian, K.; Padmanabhan, R.G.; Selvaraj, R.; Bagal, D.K.; Rajesh, S. A Review on PLA with Different Fillers
Used as a Filament in 3D Printing. Mater. Today Proc. 2022, 50, 2057–2064. [CrossRef]
15. Calì, M.; Pascoletti, G.; Gaeta, M.; Milazzo, G.; Ambu, R. New Filaments with Natural Fillers for FDM 3D Printing and Their
Applications in Biomedical Field. Procedia Manuf. 2020, 51, 698–703. [CrossRef]
16. Li, T.; Aspler, J.; Kingsland, A.; Cormier, L.M.; Zou, X. 3d Printing—A Review of Technologies, Markets, and Opportunities for
the Forest Industry. J. Sci. Technol. For. Prod. Process 2016, 5, 30.
Polymers 2022, 14, 4930 16 of 16

17. Le Duigou, A.; Castro, M.; Bevan, R.; Martin, N. 3D Printing of Wood Fibre Biocomposites: From Mechanical to Actuation
Functionality. Mater. Des. 2016, 96, 106–114. [CrossRef]
18. Azadi, M.; Dadashi, A.; Dezianian, S.; Kianifar, M.; Torkaman, S.; Chiyani, M. High-Cycle Bending Fatigue Properties of Additive-
Manufactured ABS and PLA Polymers Fabricated by Fused Deposition Modeling 3D-Printing. Forces Mech. 2021, 3, 100016. [CrossRef]
19. Wang, Z.; Xu, J.; Lu, Y.; Hu, L.; Fan, Y.; Ma, J.; Zhou, X. Preparation of 3D Printable Micro/Nanocellulose-Polylactic Acid
(MNC/PLA) Composite Wire Rods with High MNC Constitution. Ind. Crops Prod. 2017, 109, 889–896. [CrossRef]
20. Yao, T.; Deng, Z.; Zhang, K.; Li, S. A Method to Predict the Ultimate Tensile Strength of 3D Printing Polylactic Acid (PLA)
Materials with Different Printing Orientations. Compos. Part B Eng. 2019, 163, 393–402. [CrossRef]
21. Chacon, J.M.; Caminero, M.A.; Garcia-Plaza, E.; Nunez, P.J. Additive Manufacturing of PLA Structures Using Fused Deposition Mod-
elling: Effect of Process Parameters on Mechanical Properties and Their Optimal Selection. Mater. Des. 2017, 124, 143–157. [CrossRef]
22. Tanikella, N.G.; Wittbrodt, B.; Pearce, J.M. Tensile Strength of Commercial Polymer Materials for Fused Filament Fabrication 3D
Printing. Addit. Manuf. 2017, 15, 40–47. [CrossRef]
23. Mishra, R.; Wiener, J.; Novotna, J. Bio-Composites Reinforced with Natural Fibers: Comparative Analysis of Thermal, Static and
Dynamic-Mechanical Properties. Fibers Polym. 2020, 21, 619–627. [CrossRef]
24. Tiwary, V.K.; Arunkumar, P.; Kulkarni, P.M. Micro-Particle Grafted Eco-Friendly Polymer Filaments for 3D Printing Technology.
Mater. Today Proc. 2020, 28, 1980–1984. [CrossRef]
25. Kariz, M.; Sernek, M.; Obućina, M.; Kuzman, M.K. Effect of Wood Content in FDM Filament on Properties of 3D Printed Parts.
Mater. Today Commun. 2018, 14, 135–140. [CrossRef]
26. Daver, F.; Lee, K.P.M.; Brandt, M.; Shanks, R. Cork–PLA Composite Filaments for Fused Deposition Modelling. Compos. Sci.
Technol. 2018, 168, 230–237. [CrossRef]
27. Mishra, R.; Tiwari, R.; Marsalkova, M.; Behera, B.K. Effect of TiO2 Nanoparticles on Basalt/Polysiloxane Composites: Mechanical
and Thermal Characterization. J. Text. Inst. 2012, 103, 1361–1368. [CrossRef]
28. Abdullah, A.H.; Alias, S.K.; Abdan, K.; Ali, A. A study of fatigue life of kenaf fibre composites. In Advances in Manufacturing and
Materials Engineering (ICAMME); Trans Tech Publications Ltd.: Wollerau, Switzerland, 2012; Volume 576, pp. 757–760.
29. Senatov, F.S.; Niaza, K.V.; Stepashkin, A.A.; Kaloshkin, S.D. Low-Cycle Fatigue Behavior of 3d-Printed PLA-Based Porous
Scaffolds. Compos. Part B Eng. 2016, 97, 193–200. [CrossRef]
30. Essassi, K.; Rebiere, J.-L.; El Mahi, A.; Ben Souf, M.A.; Bouguecha, A.; Haddar, M. Experimental and Analytical Investigation of
the Bending Behaviour of 3D-Printed Bio-Based Sandwich Structures Composites with Auxetic Core under Cyclic Fatigue Tests.
Compos. Part A Appl. Sci. Manuf. 2020, 131, 105775. [CrossRef]
31. Gama, N.; Ferreira, A.; Barros-Timmons, A. 3D Printed Cork/Polyurethane Composite Foams. Mater. Des. 2019, 179, 107905. [CrossRef]
32. Tao, G.; Xia, Z. A Non-Contact Real-Time Strain Measurement and Control System for Multiaxial Cyclic/Fatigue Tests of Polymer
Materials by Digital Image Correlation Method. Polym. Test. 2005, 24, 844–855. [CrossRef]
33. Chen, F.; Xu, Q.; Huang, F.; Xie, Z.; Fang, H. Effect of Nozzle Vibration at Different Frequencies on Surface Structures and Tensile
Properties of PLA Parts Printed by FDM. Mater. Lett. 2022, 325, 132612. [CrossRef]
34. Magalhães da Silva, S.P.; Antunes, T.; Costa, M.E.V.; Oliveira, J.M. Cork-like Filaments for Additive Manufacturing. Addit. Manuf.
2020, 34, 101229. [CrossRef]
35. Jerez-Mesa, R.; Travieso-Rodriguez, J.A.; Llumà-Fuentes, J.; Gomez-Gras, G.; Puig, D. Fatigue Lifespan Study of PLA Parts
Obtained by Additive Manufacturing. Procedia Manuf. 2017, 13, 872–879. [CrossRef]

You might also like